oxidation of cobalt nanocrystals: investigation of the role of ......oxidation process of cobalt...

30
llLook.com Oxidation of Cobalt Nanocrystals: Investigation of the Role of Nanocrystallinity, Self-Ordering and Nanocrystal Size Master of Science Thesis in the Master Degree Program, Materials Chemistry and Nanotechnology JOHANNA BERGSTRÖM Department of Chemical and Biological Engineering Division of Applied Chemistry CHALMERS UNIVERSITY OF TECHNOLOGY Göteborg, Sweden 2013

Upload: others

Post on 26-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • llLook.com

    Oxidation of Cobalt Nanocrystals: Investigation of

    the Role of Nanocrystallinity, Self-Ordering and

    Nanocrystal Size Master of Science Thesis in the Master Degree Program,

    Materials Chemistry and Nanotechnology

    JOHANNA BERGSTRÖM

    Department of Chemical and Biological Engineering

    Division of Applied Chemistry

    CHALMERS UNIVERSITY OF TECHNOLOGY

    Göteborg, Sweden 2013

  • i

    Oxidation of Cobalt Nanocrystals:

    Investigation of the Role of Nanocrystallinity,

    Self-Ordering and Nanocrystal Size

    JOHANNA BERGSTRÖM

    The thesis work was performed at Université Pierre et Marie Curie (UPMC) in

    Paris, France in the Laboratoire des Matériaux Mésoscopiques et Nanométriques

    and as a part of the SupraNano project group. The project group is directed by

    Professor Marie-Paule Pileni.

    Department of Chemical and Biological Engineering

    CHALMERS UNIVERSITY OF TECHNOLOGY

    Göteborg, Sweden 2013

  • ii

    Oxidation of Cobalt Nanocrystals: Investigation of the Role of

    Nanocrystallinity, Self-Ordering and Nanocrystal Size

    JOHANNA BERGSTRÖM

    © JOHANNA BERGSTRÖM 2013

    Department of Chemical and Biological Engineering

    Chalmers University of Technology

    SE-412 96 Göteborg

    Sweden

    Telephone + 46 (0)31-772 1000

    Cover: HRTEM images from polycrystalline fcc nanocrystals produced from organometallic

    synthesis. One of the obtained structures after oxidation at 200˚C for 10 min, yolk/shell fcc-

    Co/CoO particle showing the lattice planes for the fcc-Co core and the CoO shell. See page 14

    for more information.

    Göteborg, Sweden 2013

  • iii

    Acknowledgements

    I would like to thank Professor Marie-Paule Pileni for taking me into her group at Laboratoire des

    Matériaux Mésoscopiques et Nanométriques, at UPMC in Paris. Thank you also to Zhijie Yang and

    Dr. Jianhui Yang for valuable help and for answering all of my questions and thank you Professor

    Krister Holmberg for putting me in contact with Professor Pileni and for valuable help from Sweden.

  • iv

    Abstract The use of cobalt and cobalt oxide nanocrystals has proven promising in technical applications such as

    information storage, catalysis and Li-ion batteries. Hollow cobalt oxide nanocrystals can be obtained

    by using the nanoscale Kirkendall effect and the ability to tune the physical properties has made

    research regarding size, shape, crystal structure and composition into a popular field. However, the

    effect of the nanocrystallinity on ordered nanoparticles exposed to oxidation has not been considered

    this extensively in previous research. The effect of nanocrystallinity has been carried out using four

    different samples, all 8 nm before oxidation, of cobalt nanocrystals synthesized by two different

    routes, the reverse micelle route for the samples small domain polycrystals and single domain hcp

    nanocrystals. Synthesized by the second route, the organometallic synthesis, were the other two

    samples, polycrystalline fcc nanocrystals and single domain ε nanocrystals. The samples were exposed

    to an oxygen flow for 10 minutes during heating at 200°C or 260°C after which the nanocrystals were

    characterized using TEM, HRTEM and ED. An investigation of how size affects the oxidation

    behavior was carried out by using 4, 6 and 8 nm samples from the reverse micelle route. Poly- and

    single domain crystal Co-nanoparticles showed different oxidation behavior. Single crystalline Co-

    nanoparticles have a higher tendency to form hollow nanoparticles than polycrystalline Co-

    nanoparticles. The difference in oxidation behavior is also clear between the two single crystalline and

    between the two poly crystalline structures. Isolated nanocrystals have a higher probability of being

    fully oxidized than ordered nanocrystals and the size of the nanocrystals as well as the oxidation

    temperature has an effect on which oxide, the cubic CoO or the spinal Co3O4, is obtained after

    oxidation.

    Keywords: Co nanoparticles, nanocrystallinity, 2D ordering effect, nanocrystal size effect, core/shell

    structures, hollow structures, yolk/shell structures.

  • v

    Table of Contents 1. Introduction ..................................................................................................................................... 1

    1.1. Purpose .................................................................................................................................... 1

    1.2. Limitations............................................................................................................................... 2

    1.3. Outline of the thesis ................................................................................................................. 2

    2. Theory ............................................................................................................................................. 3

    2.1. The Kirkendall effect ............................................................................................................... 3

    2.2. Crystal structures of metallic cobalt ........................................................................................ 3

    2.3. Structures of Cobalt oxides ..................................................................................................... 4

    2.4. Evaporation-induced self-assembly ......................................................................................... 5

    2.4.1. Colloidal stabilization by oleic acid ................................................................................ 5

    3. Method............................................................................................................................................. 6

    3.1. Experimental Procedure .......................................................................................................... 6

    3.1.1. Sample preparation for TEM ........................................................................................... 7

    3.1.2. Dry phase annealing ........................................................................................................ 7

    3.1.3. Dry phase oxidation ......................................................................................................... 7

    3.2. Characterization ....................................................................................................................... 8

    3.2.1. Transmission Electron Microscopy (TEM) ..................................................................... 8

    3.2.2. Electron diffraction (ED) ................................................................................................. 8

    3.3. Statistics................................................................................................................................... 8

    4. Results and Discussion .................................................................................................................... 9

    4.1. Nanocrystallinity dependency ................................................................................................. 9

    4.1.1. Polycrystalline nanocrystals produced from reversed micelles (Copoly-RM) ..................... 9

    4.1.2. Single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse micelle approach (CoHCP-RM) .......................................................................................... 11

    4.1.3. Polycrystalline fcc nanocrystals produced from organometallic synthesis (Copoly-ORG) 14

    4.1.4. Single domain ε phase Co nanocrystals produced from organometallic synthesis (Coε-ORG) ................................................................................................................................. 16

    4.2. Size dependency .................................................................................................................... 18

    4.2.1. Polycrystalline nanocrystals produced from reverse micelles (Copoly-RM) ..................... 18

    4.2.2. Single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse micelle approach (CoHCP-RM) .......................................................................................... 20

    5. Conclusion ..................................................................................................................................... 22

    5.1. Future work ........................................................................................................................... 23

  • 1

    1. Introduction

    Metal oxides are commonly used today in a variety of applications from cathodes in batteries to

    semiconductors in solar cells (1), (2). Hollow inorganic nanocrystals possess the interesting

    phenomena of having an empty internal space which creates possibilities for usage in optical,

    catalytically and magnetic devices such as nanoreactors and carriers for drug delivery (3). The ability

    to tune physical properties of metal oxides nanoparticles has done the investigation regarding size,

    shape, crystal structure and composition into a popular field (4), (3). However, not as much attention

    has been focused on how the crystallinity of the nanoparticle, called the nanocrystallinity, affects the

    physical properties (4), (5).

    Magnetic transition metals, like cobalt (Co), are often air sensitive which makes them complicated to

    work with. As a consequence of this, metal oxides are used in magnetic applications instead despite

    the weaker magnetic properties, as they are stable in an air containing environment (6). Co and cobalt

    oxides (CoO, Co3O4) are promising materials when it comes to usage in technical applications such as

    catalysis, information storage and Li-ion batteries (7). CoO hollow nanocrystals can be successfully

    synthesize by the Kirkendall effect starting from solid Co nanoparticles and exposing them to a flow

    of oxygen (8). The Kirkendall phenomenon of diffusion arises during heating when two species in a

    diffusion couple have different atomic diffusivities (9). As an effect, pores will form in the faster

    diffusing material, because of a motion of the boundary layer, and can be supersaturated to form a

    single interior void or a yolk-shell structure (8), (10). Diffusion is made easier by defects in the lattice,

    like grain boundaries, since this enables atomic jumps. A hollow core formation can be prevented by

    the produced phase being rich in defects which can absorb the vacancies without forming voids or by

    the particle being too small to be stable as a hollow structure (11). The synthesis of hollow

    nanocrystals through the Kirkendall effect provides a template free route which automatically

    eliminates the sometimes difficult step of removing the template before obtaining the hollow particle

    (3).

    It has been shown that the nanocrystallinity of isolated 7 nm Cobalt nanocrystals does not play a role

    for oxidation, as they are completely oxidized independently of their nanocrystallinity. As the particles

    are assembled into a compact hexagonal network in a 2D lattice, the oxygen diffusion rate will be

    slowed down, and the oxidation will be affected by the nanocrystallinity, leading to either core/shell

    Co/CoO particles or hollow particles of CoO. (7) The effect of the nanocrystallinity on ordered

    nanoparticles exposed to oxidation has not been considered this extensively in previous research.

    1.1. Purpose

    The purpose of the thesis is to investigate the control of nanocrystallinity through analysis of the

    oxidation process of cobalt nanocrystals upon change in the crystalline structure of the nanoparticles

    and their assemblies, and also to examine how nanocrystal size affects the oxidation process.

    Investigations of four different cobalt samples have been made, fcc and hcp synthesized by reverse

    micelle route, and fcc and ε formed by organometallic synthesis. The difference between the two fcc

    samples, synthesized by different the routes, is that the reverse micelle route results in the formation of

    smaller crystalline domains and thereby a more poorly crystallized structure, called small domain

    polycrystal, than the organometallic synthesis. The two fcc structures therefor differ in

    nanocrystallinity, as their domain size differs and the one with smaller domain size is closer to an

    amorphous structure, even if they are both fcc and polycrystalline.

  • 2

    1.2. Limitations

    Oxidation, annealing and characterization are a part of the thesis work however synthesis of the Co-

    nanoparticles will not be included, but provided by the lab. For this reason, no synthetic routes will be

    presented as the focus has been on the characterization and of the nanoparticles. Metallic cobalt has

    three different crystal structures, which are all described more thoroughly in chapter 2.2, and the three

    possible structures have been tested.

    Nanocrystallinity dependency will be investigated using nanoparticles with a size of 8 nm and how

    size affect the oxidation process will be examined for three different sizes, 4, 6 and 8 nm and only for

    two of the samples, the crystal structures fcc and hcp synthesized by the reverse micelle route. The

    reason why size dependency is not performed for all the samples is that there is not yet a successful

    method to obtain a good enough size control for the smaller particles made by the organometallic

    synthesis.

    HRTEM has been performed on the samples before oxidation and for the samples oxidized at 200°C

    for 10 minutes but not for the samples oxidized at 260°C for the same time. The reason for this is that

    the HRTEM is a time consuming process and the instrument can only be run by trained people. In the

    end the time was not enough and the 200°C oxidation treatment was prioritized.

    1.3. Outline of the thesis

    The report is divided into five chapters, chapter 2, Theory, is presenting the necessary background

    information needed to understand the experimental methods and the oxidation behavior. It deals with

    the diffusion behavior described by the Kirkendall effect, the different structures of cobalt and cobalt

    oxides, the self-assembly and the colloidal stabilization necessary to prevent the particles from

    coalescence. Chapter 3 deals with the methods which in more elaborated terms mean the experimental

    procedure and the description of the methods for characterization, Transmission electron microscopy

    (TEM), High Resolution TEM (HRTEM) and electron diffraction (ED). Chapter 4, Results and

    Discussion, is divided into two parts. The first part handles the effect of nanocrystallinity and self-

    ordering while the second part presents the results regarding the size effect. The thesis is ended with

    chapter 5, conclusions, concluding the work and suggesting further areas of investigations.

  • 3

    2. Theory

    The theory chapter contains background about the Kirkendall effect used to create hollow CoO or

    core/shell Co/CoO, information about the three different crystal structures of metallic cobalt and

    cobalt oxide, and the theory behind the self-assembly used to deposit samples on TEM grids for

    characterization.

    2.1. The Kirkendall effect

    During solid state reactions, diffusivity is a kinetically determining factor. Charge and size of the

    particles taking place in the reaction control the diffusivity. If two particles with different diffusivity

    react with each other the reaction will be faster in one direction than the other. This is known as the

    Kirkendall effect and results in a motion of the boundary layer that can be used to produce hollow

    nanoparticles. Oxygen is a large and charged ion and even though the metal ions in the lattice are also

    charged, the smaller size of the metal ion compared to the oxygen is what most often make up for the

    diffusion difference. This causes the metal ions to migrate outwards rather than the oxygen ions to

    migrate into the metal lattice, the created vacancies migrate inwards and can form a single interior

    cavity. (10)

    Figure 1. The nanoscale Kirkendall effect. The metal nanocrystal is reacted with oxygen and because of the faster

    outward diffusion of metal than the inward diffusion of oxygen voids are formed in the lattice. The voids are saturated

    in the center leading to a hollow nanocrystal.

    The Kirkendall effect is known for causing weakness within a bulk material as it is the reason for

    internal void formation and thereby weakening of the material, but on the nanometer scale it can be

    used to create hollow nanoparticles instead, because of super saturation of vacancies into a hollow

    core (3), (11)., see Figure 1. The nanoscale Kirkendall effect is more complex than on a macroscale

    since the nanoscale system include factors like a shorter diffusion length, interface stress and

    inhomogeneous coating, all of which are less significant on a macroscale. Diffusion is made easier by

    defects in the lattice, like grain boundaries, since this enables atomic jumps. A hollow core formation

    can be prevented the produced phase being rich in defects which can absorb the vacancies without

    forming voids or by the particle being too small to be stable as a hollow structure. (11)

    2.2. Crystal structures of metallic cobalt

    Cobalt is a transition metal and that can be found between iron and nickel in the periodic system of

    elements and it is naturally occurring in the crust of the earth (2). Metallic cobalt has three possible

    crystal structures, the hexagonal closed-packed (hcp), face-centered cubic (fcc) and epsilon phase (ε)

    (12). The three phases are shown in in Figure 2 A-C. At normal pressure, two crystal structures are

    stable for bulk Co, the hcp below 425°C and the fcc at higher temperatures (2), (13). Because of a low

    free energy exchange, the change into the more stable hcp in room temperature is very slow (2).

  • 4

    Metastable phases such as ε-Co nanocrystals are possible to form in solution phase chemical synthesis

    because it is as a general rule not controlled by thermodynamics.

    The hcp structure is constructed of layers in an ABABAB-stacking sequence, this leaves small

    channels in the structure between the Co-atoms. The fcc structure has an ABCABC-stacking sequence,

    making it cubic. The ε-phase is composed of two different geometrical arrangements, having the cubic

    structure of β-manganese, some distances are shorter than others and the purple and blue atoms in

    Figure 2 C have their neighboring atoms arranged differently.

    Figure 2. Sketch view of different bulk cobalt structures: (A) hcp or α, (B) fcc or β, (C) ε, (including a detail of the

    different cobalt position). [From reference (12), Reprinted with permission from American Institute of Physics. ©

    2010 American Institute of Physics.]

    During annealing, heat treatment, of a Co sample at a temperature below 425 ᵒC the metal ions in the

    lattice becomes more mobile than in room temperature. This causes the ε or fcc lattice to transform

    into the more stable hcp structure (6), (14). The mechanics behind the change is due to dislocation

    movements of the octahedral planes of the cubic fcc or ε lattice (2).

    Single domain crystals are built up by one single crystal, while polycrystals are built up by two or

    more fused crystals which mean that grain boundaries are present within a polycrystal. Magnetic

    properties of nanocrystals depend strongly on both particle size and the crystal structure. The complex

    cubic structure of the ε-Co nanocrystal is a soft magnetic material meaning the coercivity is low and

    the required force to demagnetize the material after it has been magnetized is low (6). The hcp is

    ferromagnetic at all temperatures, and the magnetic behavior remains after the source of the alignment

    of the electric dipoles has been removed. However the fcc becomes paramagnetic at 1121 ± 3ᵒC,

    resulting in the consequence that the electric dipoles orient randomly after the removal of the magnetic

    field above the given temperature. Single crystalline hcp is magnetically anisotropic and will be more

    easily magnetized in on direction than the other, while the cubic structures have an isotropic magnetic

    behavior. The magnetic behavior of polycrystalline particles depend both of the purity of the metal and

    the thermal history of the material (2).

    2.3. Structures of Cobalt oxides

    Cobalt oxide complexes have been traditionally used as dying agents in the ceramic and glass

    production, it adapts different color depending on the metal ion binding together with the cobalt oxide.

    The density for the oxides is lower than for metallic cobalt leading to a growth in size of cobalt

    nanoparticles during oxidation (2).

  • 5

    Cobalt has two readily available oxidation states, Co2+

    and Co 3+

    , and cobalt oxides exist in two main

    structures. The structurally most simple one is cobalt(II)oxide (CoO) which has a rock salt structure,

    meaning one Co2+

    is octahedrally coordinated to the oxygen atoms in the lattice, with a = 0.425 nm.

    The structure is stable above 900 ᵒC and is antiferromagnetic at normal temperatures. CoO has a

    density of 6.44 g/cm3

    at 20ᵒC (15).The second main structure is cobalt(II,III)oxide (Co3O4), it is

    divalent and has a spinel structure with tetrahedrally coordinated Co2+

    and octahedrally coordinated

    Co3+

    , with a = 0.807 nm. It is the thermodynamically stable form of cobalt oxide under ambient

    temperature and pressure (1), (2), (16). The larger separation between the lattice fringes means that

    Co3O4 has a lower density than CoO with a value of 6.11g/cm3 at 20ᵒC

    (15).

    Cobalt oxide exists in a third composition as well, the cobalt(III)structure (Co2O3), it is unstable but

    impure forms have been prepared. There is never a transition from Co3O4 into Co2O3, but the

    chemistry of Co(III) is instead dominated by formation of other chemical complexes (1), (2), (16).

    The oxides can exist as either poly- or single crystalline structures which will affect their properties

    (2). Polycrystals can be used in catalysts or as nanoscale reactors as the grain boundaries enable a

    possible route for diffusivity of small particles (8).

    2.4. Evaporation-induced self-assembly

    The compact hexagonal 2D networks studied in the project are self-assembled through evaporation-

    induced self-assembly. A stable colloidal dispersion of Co nanocrystals is deposited on a surface and

    the solvent is allowed to evaporate. The ordering takes place in the meniscus, which is the curved

    surface formed at the top of the liquid, and because of the small volume of it, the nanocrystals are

    forced together. The two most important repulsive interactions of the system are the steric hindrance

    from the surfactants covering the surface of each nanocrystal due to reduced entropy and the osmotic

    flow of solvent towards the compressed region, and the columbic repulsions due to the same signs of

    the surface charges. The two forces together maximize the distance between the nanocrystals which

    creates an ordered compact network. (10)

    2.4.1. Colloidal stabilization by oleic acid

    All samples presented in the report are coated with oleic acid, C18H34O2, see Figure 3. Steric

    stabilization is necessary to avoid aggregation in the colloidal dispersion and the carboxylic group in

    oleic acid binds strongly to the particle surface and forces the long hydrocarbon chain out into the non-

    polar solution (17). The long hydrocarbon chain probably also has an effect on the oxygen diffusion as

    is will work as a barrier for the nanocrystal and a protection against oxidation as well as aggregation.

    Oleic acid has a melting point at 13 – 14ᵒC and a boiling point at 194 – 195ᵒC (18).

    Figure 3. Chemical structure of oleic acid.

    H3C (CH2)7 CH CH (CH2)7 C OH

    O

  • 6

    3. Method

    The chapter describes the experimental procedure, the characterization methods and statistics. The

    experimental procedure starts with an overview before going into the different steps, after which

    characterization methods are presented one by one before a short section on how statistical certainty

    was achieved during analysis.

    3.1. Experimental Procedure

    Investigation of four different nanocrystalline cobalt samples have been performed, fcc synthesized by

    reverse micelles, which is a polycrystal with small crystalline domains (sample 1), hcp synthesized by

    reversed micelles, which has a single crystalline structure (sample 2), fcc formed by organometallic

    synthesis, which is a polycrystal (sample 3) and ε formed by organometallic synthesis which has a

    single crystalline structure (sample 4). The four investigated samples are summarized in Table 1

    below. The difference between the two fcc samples, synthesized by the different routes, is that the

    reverse micelle route results in the formation of smaller crystalline domains and thereby a more poorly

    crystallized structure, called small domain polycrystal, than the organometallic synthesis. The two fcc

    structures therefor differ in nanocrystallinity, as their domain size differs and the one with smaller

    domain size is closer to an amorphous structure, even if they both have the fcc structure and are

    polycrystalline.

    Table 1. Summary of investigated samples, their structure, nanocrystallinity and the synthetic route for formation.

    Sample no. Structure Nanocrystallinity Synthetic route Size (nm) Abbreviation

    1 fcc Small domain polycrystal Reverse micelle 4, 6, 8 Copoly-RM

    2 hcp Single domain Reverse micelle 4, 6, 8 CoHCP-RM

    3 fcc Polycrystal Organometallic

    synthesis

    8 Copoly-ORG

    4 ε Single domain Organometallic

    synthesis

    8 Coε-ORG

    The stabilizer of the colloidal dispersion was in all four cases oleic acid, which is described in chapter

    2.4.1. Comparison of the nanocrystallinity effect was performed with nanoparticles in the size of 8 nm

    and the size effect has been examined using three different sizes, 4, 6 and 8 nm, of fcc and hcp

    synthesized by the reverse micelle route. Characterization of the nanoparticles was executed using

    TEM, HRTEM and electron diffraction.

    Sample 1, 3 and 4 were synthesized in a glove box with nitrogen atmosphere to avoid oxidation of the

    nanoparticles. To achieve the hcp structure of sample 2, sample 1 was dry-phase annealed in 250 ᵒC

    under argon flow for 60 minutes. All samples were oxidized in 200˚C or 260˚C for ten minutes and

    characterized with TEM and ED both before and after oxidations to investigate how the oxidation

    procedure had been affected by the nanocrystallinity, the self-ordering and the nanocrystal size.

    HRTEM was performed on all samples after oxidation to receive more information about the oxidation

    behavior. A summary of the experimental procedure is presented in the flowchart in Figure 4.

  • 7

    Figure 4. Flow chart of experimental procedure for investigation of how nanocrystallinity, self-ordering and

    nanocrystal size affects the oxidation of Co nanocrystals.

    3.1.1. Sample preparation for TEM

    The sample was prepared in a glove box with inert nitrogen atmosphere. The sample, containing Co-

    nanoparticles coated with oleic acid in hexane, was deposited on the TEM grid (copper grid coated

    with amorphous carbon film) and left for two minutes allowing the solvent to evaporate causing the

    Co-nanoparticles to self-assemble into a monolayer. The grid was placed in the sample loader and

    inserted into the TEM for characterization.

    3.1.2. Dry phase annealing

    The samples were annealed through dry phase annealing, meaning that the particles were already

    deposited and ordered into a 2D lattice on the TEM grid when exposed to the heat treatment. The grid

    with the applied sample was placed in a tube and loaded on a modified Schlenk line, which included a

    high vacuum pump, a source of inert argon gas and a source of oxygen gas. The sample was exposed

    to vacuum then Ar flow in three rounds to ensure that all air had been removed from the system. The

    heater was adjusted to 250 ᵒC and the annealing took place under argon flux for 60 minutes, after

    which the heater was removed to stop the annealing. The sample was left to cool to room temperature

    by the argon flow (10 min).

    3.1.3. Dry phase oxidation

    The grid with sample on it was placed in a tube and loaded on the modified Schlenk line, the same

    used during annealing, and treated with O2 (200˚C, 10 min or 260˚C, 10 min). After oxygen treatment

    the heater was removed, the oxygen flow was replaced by argon flow to stop the oxidation

    immediately and the sample was allowed to cool to room temperature by the argon flow (10 min).

  • 8

    3.2. Characterization

    Methods used for characterization of the Co-nanoparticles, both before and after the two oxidations,

    are TEM, HRTEM and ED. The methods are described in the two following subchapters.

    3.2.1. Transmission Electron Microscopy (TEM)

    The analysis is performed in a vacuum chamber. A high energy beam of electrons is directed towards

    a thin sample of the particles assembled on a copper grid coated with an amorphous carbon film. The

    electrons that are transferred through the sample are detected and form a two dimensional projection of

    the sample. The instrument can be chosen to detect either the direct beam (bright field image) or the

    diffracted beam (dark field image) (19). TEM can be used as a high resolution instrument to detect the

    size and shape of the particles, but not to determine the 3D structure of the lattice.

    TEM is available in different modes, depending on what instrument is used. In High Resolution mode

    (HRTEM), the electrons are accelerated with a higher potential field, giving them a shorter wavelength

    and higher energy which increase the resolution of the obtained image (19). This way the lattice planes

    of a nanocrystal oriented in the right direction can be seen.

    TEM was performed using a JEOL 1011 microscope at 100 kV, and HRTEM was performed using a

    JEOL 2010 microscope at 200 kV.

    3.2.2. Electron diffraction (ED)

    Electron diffraction (ED) is available in the TEM instrument and can be used to determine the crystal

    structures of the nanocrystals. Electrons are fired at the sample and the resulting interference pattern is

    observed as the incoming primary electrons interact with the secondary electrons reflected off the

    lattice planes. Electron diffraction can be used to calculate the interplanar distance between lattice

    planes and can thereby be used to determine the crystal structure (20). The diameter of the resulting

    rings in the ED patterns are measures and the inverse value of the radius can be compared to

    theoretical values to determine, confirm or deny the crystal structure.

    3.3. Statistics

    To measure size of the full diameter or the internal structure and achieve histograms and size

    distributions of the nanocrystals both before and after oxidation, at least 500 nanocrystals were

    measured on the TEM images for each sample, using the free software ImageJ. The histograms were

    prepared using Origin8. In the cases where the percentage of hollow, core/shell or fully oxidized

    particles are presented, at least 500 nanocrystals were investigated, using the TEM images.

  • 9

    4. Results and Discussion

    The chapter is divided into two main subchapters, the first one, chapter 4.1, concerns the investigation

    on how the nanocrystallinity affects the oxidation behavior and the second one, chapter 4.2, is about

    how the size of the nanoparticle affects the oxidation behavior. Both subchapters deal with the

    question on how the self-ordering affects the end results. The nanocrystallinity dependency is

    examined using 8 nm particles for all four different samples while the size effect is evaluated using 4,

    6 and 8 nm particles of Copoly-RM and CoHCP-RM. The expected result before the start of the investigation

    was to find a change within the oxidation of the various single domains and between the single- and

    polycrystals.

    4.1. Nanocrystallinity dependency

    The results regarding the 8 nm particles for each of the four different nanocrystallinities are presented

    separately for each sample and discussed in the text. The samples are all presented before oxidation,

    after oxidation at 200°C for 10 minutes and after oxidation at 260°C for 10 minutes. After the last

    sample a more general discussion is made, taking all the results into account and trying to answer how

    nanocrystallinity and self-ordering affects the oxidation behavior of cobalt nanocrystals.

    4.1.1. Polycrystalline nanocrystals produced from reversed micelles (Copoly-RM)

    The polycrystalline nanocrystals synthetized by the reverse micelle route (Copoly-RM) are characterized

    by very small fcc domains that indicate a very low crystallinity. No evidence can be given to prove

    that some domains are not amorphous. In Figure 5a, a Copoly-RM nanoparticle is visualized in a HRTEM

    image. No crystalline domains are clearly seen as they are too small to appear as an ordered structure

    in the images, this indicates that the crystallinity is not well defined.

    Figure 5. HRTEM images from polycrystalline nanocrystals produced from reversed micelles (Copoly-RM) 8 nm.

    (a) Native nanoparticle before oxidation, no lattice planes can be perceived which indicates low crystallinity, (b) After

    oxidation at 200˚C for 10 min. Core/shell Copoly-RM/CoO nanoparticle showing the lattice plane (200) for CoO in the

    shell.

    Figure 6a and 6b show the TEM images at various scales of the nanocrystals in their native state

    before oxidation. The observed zone is well ordered in a compact hexagonal network with some

    isolated nanocrystals, giving the opportunity to study how the 2D ordering affects the oxidation. From

    the histogram (inset Figure 6a) the nanocrystal size and its distribution are 8.3 ± 0.77 nm and 9.3%

    respectively. The electron diffraction (ED) pattern (Figure 6c) shows only one clear ring

    corresponding to the fcc lattice plane (111). This confirms the information from the HRTEM image of

    the same sample in Figure 5a in which the crystallinity of the nanoparticles is not well defined.

  • 10

    Figure 6. Polycrystalline nanocrystals produced from reversed micelles (Copoly-RM), 8 nm. (a, b) TEM images before

    oxidation, (d, e) after oxidation at 200˚C for 10 min, (g, h) after oxidation at 260˚C for 10 min. (c, f, i) Corresponding

    ED pattern before and after the two oxidation temperatures. The insets in (a, g) are histograms displaying the size

    distributions of the nanoparticle diameter, in (d) the histogram displays both the particle diameter and the diameter

    of the core in the core/shell nanoparticle.

    The TEM images of the sample submitted to oxidation at 200°C for 10 minutes reveals formation of

    both core/shell (Copoly-RM/CoO) and CoO nanocrystals (Figure 6d and 6e). The average diameter of

    nanocrystals is 10.6 ± 0.98 nm, with a size distribution of 9.2%, which is almost equal to the

    distribution observed before oxidation (9.3%) even if the nanoparticles have grown. The average

    diameter of the core proved to be 3.9 ± 1.07 nm, with a size distribution of 27.8%. The calculated

    thickness of the oxide shell is thereby 3.4 nm, however, there is large size distribution of the core

    making the calculations uncertain. From HRTEM image (Figure 5b) it can be seen that the crystalline

    structure of Co core remains similar as that observed before oxidation while the oxidized shell shows a

    polycrystalline structure with a lattice plane corresponding to the CoO (200). A particle in more highly

    ordered areas of the lattice has the tendency to become core/shell during oxidation, but along the edges

    of the monolayer and in the regions where the nanocrystals are not well ordered, most of the

    nanocrystals are fully oxidized.

    In Figure 7a, a small assembly of nanoparticles is visualized, it is clearly seen that the nanoparticles in

    the middle of the assembly are core/shell while the edges of the assembly consists of nanoparticles

    with a smaller core or are fully oxidized. It can be noticed that isolated nanoparticles are mostly fully

    oxidized, but some size dependency can be seen among them, where larger particles received a

    core/shell structure while smaller particles were fully oxidized. In Figure 7b two isolated particles

    show a core/shell structure with a very small core. A general feature is that the larger the nanoparticle

    size is, the larger the core becomes in the core/shell structure for both ordered and isolated structures.

    The size effect of the Copoly-RM has been further investigated in chapter 4.2.1. The ED pattern (Figure

    6f) displays rings for both CoO and fcc-Co supporting the information given by the HRTEM image

  • 11

    that a core shell structure with both the fcc-Co and CoO is present. However, only the outer rings of

    the fcc lattice can be seen but as fcc (111) and (220) is close in d-spacing to CoO (200) and (311) they

    can be hidden behind each other, or the rings for the fcc lattice can also be unclear because of the

    nearly amorphous structure. The separation between the lattice planes is larger for CoO than fcc-Co,

    and this change in density explains the growth during oxidation (2), (15).

    Figure 7. Polycrystalline nanocrystals produced from reversed micelles (Copoly-RM), 8 nm, after oxidation at 200˚C for

    10 min. (a) TEM images of small assembly of nanoparticles. (b) TEM image of isolated particles. The inset is the

    corresponding low magnification TEM where the enlarged image is selected from the white rectangle.

    When the oxidation process takes place at 260°C instead of 200°C, keeping the same exposure time

    (10 min) some core/shell structure is still present, but most of the nanoparticles have become fully

    oxidized (Figure 6g and 6h). The size of the particles has increased to 11.7 ± 1.09 nm but keeping the

    distribution of 9.3%. According to the corresponding ED pattern (Figure 6i) the oxide has obtained the

    spinal structure, Co3O4, in contradiction with the oxidation at 200°C where the nanoparticles received

    the cubic CoO structure. The space between the nanoparticles decreases on increasing the temperature

    and consequently the tendency of nanocrystals to coalescence increases. This can be due to the faster

    oxidation at 260˚C compared to 200˚C or a consequence of the smaller space between the lattice

    planes and thereby lower density of the Co3O4 compared to the CoO (15).

    In conclusion, Copoly-RM nanocrystals produced from the reverse micelle route is highly polycrystalline

    and thereby presents many grain boundaries for possible oxygen diffusion. As the oxygen diffusion

    becomes fast, the inward diffusion can compete with the outward diffusion of Co atoms and no

    internal void is formed. The oxidation thereby results in a core/shell particle. In the case that internal

    pores are formed they can be annihilated in the high concentration of grain boundaries without

    forming an internal void (11). In the less ordered areas the Copoly-RM nanoparticles are exposed to more

    oxygen compared to the nanoparticles in the more highly ordered areas, as the nanoparticles are not

    being protected from the oxygen by the neighboring particles on the sides, resulting in fully oxidized

    nanoparticles instead of the core/shell particle. These data are in agreement compared to those already

    obtained with 7nm Copoly-RM nanoparticles performed earlier by the same group (7). The higher

    oxidation temperature of 260˚C speeds up the oxidation, which results in a higher concentration of

    fully oxidized particles compared to after the lower oxidation temperature 200˚C.

    4.1.2. Single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse micelle approach (CoHCP-RM)

    Single domain crystalline hcp (CoHCP-RM) are produced by annealing Copoly-RM nanoparticles. Figure 8a

    and 8b show TEM images with well-defined nanocrystals with an average size and distribution of 8.4

    ± 0.85 nm and 10.1% respectively. The size is slightly larger than what was observed for Copoly-RM

  • 12

    nanoparticles (8.3 ± 0.77 nm and 9.3%), but the small change is within the margin of error of the

    measurements. Figure 9a shows the HRTEM images for CoHCP-RM before oxidation, the ordered lattice

    planes correspond to the hcp-Co (002) and the nanocrystal is built up by one single crystalline domain.

    No sign of oxidation from the annealing treatment can be seen as this would have shown on the

    HRTEM image and also have generated a ring, characteristic for CoO, inside the inner triplet rings in

    the ED pattern (Figure 8c).

    Figure 8. Single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse micelle approach

    (CoHCP-RM) 8 nm. (a, b) TEM images before oxidation, (d, e) after oxidation at 200˚C for 10 min, (g, h) after oxidation

    at 260˚C for 10 min. (c, f, i) Corresponding ED pattern before and after the two oxidation temperatures. Tint and Text

    marked on the ED pattern in (c) and (f) refer respectively to the internal triplet rings [(100), (101), (002)] and external

    triplet rings [(110), (103), (112)], of the hcp lattice. The insets in (a, g) are histograms displaying the size distributions

    of the nanoparticle diameter, in (d) the histogram inset displays both the particle diameter and the diameter of the

    core in the core/shell nanoparticle.

    The CoHCP-RM nanocrystals are submitted to same treatment as the previously described nanoparticles

    (oxidation at 200°C; 10mn). For any ordered nanocrystal a core/shell CoHCP-RM/CoO structure is

    observed (Figure 8d and 8e), but for isolated nanocrystals either core/shell or fully oxidized

    nanocrystals are produced. The thickness of the shell is thinner when the nanocrystals are ordered

    compared to when they are isolated, this behavior is visualized in Figure 8e with the single isolated

    particle having a smaller core, and thereby a thicker shell, than the neighboring ordered particles.

    Furthermore, the shell is thinner for the largest nanocrystals compared to smaller one when both are

    ordered in compact hexagonal network. The nanocrystals have a mean diameter of 9.4 ± 0.88 nm and a

    size distribution of 9.4%, the average diameter of the core for ordered nanoparticles is 5.8 ± 0.91 and

    has a distribution of 15.7% this gives the average shell thickness of 1.8nm. The HRTEM image

    (Figure 9b) shows that the large core keeps its former crystalline structure. It has a regular pattern of

    lattice planes that correspond to the hcp-Co (002) plane. The oxide shell shows two different distances

    for the CoO lattice, 2.13 Å corresponding to the (200) plane and 2.46 Å corresponding to the (111).

    The corresponding ED pattern (Figure 8f) illustrates a mixture of both hcp-Co and CoO which

    supports the information from the TEM and HRTEM pictures of a core/shell CoHCP-RM/CoO structure.

    These data are reproducible and surprising if compared to a previous study performed by the same

    group with 7 nm CoHCP-RM nanoparticles instead of 8 nm. For the 7 nm nanoparticles, hollow CoO

  • 13

    single domain nanocrystals were produced (7). The effect of size on the oxidation behavior of CoHCP-

    RM has been investigated further in chapter 4.2.2.

    Figure 9. HRTEM images from single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse

    micelle approach (CoHCP-RM) 8nm. (a) Before oxidation, single domain nanoparticle showing the lattice planes for the

    hcp (002) plane. (b) After oxidation at 200˚C for 10 min. Core/shell Co/CoO nanoparticle showing the lattice planes

    for the hcp core (002) and the CoO shell (111) and (200).

    When the temperature is increased to 260°C instead of 200°C, keeping the same exposure time, the

    nanoparticles tend to coalesce (Figure 8g and 8h). However, it is shown that the nanoparticles present

    hollow structure with a thick oxide shell and an internal void. An increase in the size can be seen after

    oxidation at 260˚C compared to after 200˚C as the particles are now 11.4 ± 1.33 nm with a distribution

    of 11.7%. The ED pattern (Figure 8i) shows formation of the spinal structure Co3O4 instead of CoO.

    The difference in sizes observed between the sample oxidized in 200°C and in 260°C is explained by

    the fact that Co3O4 has lower density compared to CoO and also that a more rapid oxidation and

    consequently a faster oxygen diffusion takes place after increasing the temperature. The creation of an

    internal void is also a factor that leads to a larger increase in the outer diameter

    During the annealing process it is likely that some of the oleic acid used as a coating agent is broken

    down, but as no fusion of the Co can be seen after the annealing, some coating must still be intact

    surrounding the nanoparticles and making them avoid aggregation. The partly broken oleic acid can be

    the reason that CoHCP-RM oxidized at 260°C for 10 min coalescence more than Copoly-RM after the same

    oxidation treatment. However, the formation of an internal void for CoHCP-RM and thereby a larger

    increase in the size is also a reason for enhanced coalescence. It can clearly be seen that 200°C for 10

    minutes does not produce a fast enough outward movement of cobalt compared to the inward

    movement of oxygen to create hollow nanoparticle for the 8 nm structure.

    If CoHCP-RM is compared to Copoly-RM (chapter 4.1) for the 8 nm nanoparticles, they differ in their

    oxidation behavior. Both samples form core/shell structures after oxidation at 200°C for 10 min but

    the size of the cores are larger for the hcp single crystalline structure than the polycrystalline

    nanoparticle, and the size change during oxidation is smaller for the single domain nanocrystal. The

    difference can clearly be seen when comparing the thickness of the oxide shell. For Copoly-RM the shell

    thickness is 3.9 nm while CoHCP-RM displays a thickness of 1.8 nm. The difference can be an indication

    that oxygen atoms have a slower diffusion path in to the hcp single crystal lattice than the

    polycrystalline lattice. Because of the low crystallinity many grain boundaries are present in the

    polycrystalline lattice while no grain boundaries are present in the single crystalline hcp lattice. The

    grain boundaries work as a fast diffusion path making the oxidation of easier for Copoly-RM than for

    CoHCP-RM. A faster oxidation means a smaller core for the structure after oxidation and thereby a

    thicker oxide shell.

  • 14

    4.1.3. Polycrystalline fcc nanocrystals produced from organometallic synthesis (Copoly-ORG)

    The organometallic method is used to synthesize polycrystalline fcc-Co (Copoly-ORG) with larger

    crystalline domains compared to those produced within the Copoly-RM nanocrystals, grain boundaries

    are thereby present, but not in as high concentration as for the sample prepared by reversed micelles.

    HRTEM images (Figure 10a) shows some grain boundaries, verifying the poly crystallinity of the

    sample, but if compared to Copoly-RM in Figure 5a, Copoly-ORG clearly has larger crystalline domains. The

    Copoly-ORG nanocrystals deposited on TEM grid shows mainly 2D super lattices and the formation of

    multilayers (Figure 11a and 11b) and very few isolated nanocrystals are detected. From histogram

    shown in the inset in Figure 7a, the average diameter of Copoly-ORG nanocrystals before oxidation is 8.1

    ± 0.67 nm with a 7.1% size distribution. The ED pattern (Figure 11c) clearly shows presence of an fcc

    structure.

    Figure 10. HRTEM images from polycrystalline fcc nanocrystals produced from organometallic synthesis (Copoly-ORG)

    8nm. (a) Before oxidation, lattice planes nanoparticle showing lattice planes for fcc-Co. (b) One of the obtained

    structures after oxidation at 200˚C for 10 min, Core/shell Co/CoO particle showing the lattice planes for the fcc-Co

    core and the CoO shell, (c) the other obtained structure after oxidation at 200˚C for 10 min, hollow particle showing

    the lattice planes of the CoO shell.

    After submitting Copoly-ORG nanocrystals deposited on a TEM grid to oxidation (200°C; 10 min) both

    yolk/shell (Co-fcc/CoO), and hollow CoO nanoparticles were produced (Figure 11d and 10e). With

    the term yolk/shell means that the core is isolated from the shell and that the void is saturated between

    the core and the shell unlike a core/shell particle where the metal core and the oxide shell are in

    contact with each other. The average size of nanocrystals markedly increased compared to Copoly-ORG

    nanocrystals before oxidation. It is found equal to 10.7 ± 0.80 nm with a size distribution of 7.5%

    (inset Figure 11d). The HRTEM image shows a yolk/shell nanocrystal (Figure 10b) with the Co core

    isolated from the CoO shell by voids. The core shows a single domain crystalline structure with

    regular lattice planes corresponding to the fcc-Co (111) plane. The shell shows planes with a distance

    of 2.13 Å corresponding to the CoO (200) plane. It is not possible to know if CoO shell is single

    domain or polycrystals because the orientation of the shell could be such that only one side of the shell

    shows the lattice planes, and none of the two possibilities could be confirmed or ruled out by a careful

    HRTEM study. Figure 10c shows the formation of a hollow nanocrystal, after the same oxidation

    treatment, with a shell characterized by lattice planes with a distance of 2.46 Å corresponding to the

    CoO (111) plane. The oxide shell of the hollow particle is polycrystalline, as the directions of the

    (111) planes differ in different areas. From TEM images, under investigation of 500 nanoparticles in

    monolayers 62 % of nanocrystals are characterized by a yolk/shell structure and 38 % by hollow

    nanoparticles. The shape of the particles is more irregular after oxidation than before. The

    corresponding ED pattern (Figure 11f) supports the fact that both CoO and fcc-Co are present after

    oxidation.

  • 15

    Figure 11. Polycrystalline fcc nanocrystals produced from organometallic synthesis (Copoly-ORG) 8 nm. (a, b) TEM

    images before oxidation, (d, e) after oxidation at 200˚C for 10 min, (g, h) after oxidation at 260˚C for 10 min. (c, f, i)

    Corresponding ED pattern before and after the two oxidation temperatures. The insets in (a, d, g) are histograms

    displaying the size distribution at the different times.

    The data after oxidation at 200°C for 10 min are rather surprising if compared to those obtained with

    same size of Co nanocrystals produced from reverse micelles (Copoly-RM) with those produced from

    inorganic method (Copoly-ORG). In both cases fcc domains are present before oxidation, however the

    Copoly-RM nanocrystals (Figure 5a) are characterized by small well-defined domains whereas with

    Copoly-ORG nanocrystals the fcc domains are larger (Figure 10a). Furthermore, it is observed that the

    size of crystalline fcc domains in the Copoly-ORG nanocrystals before oxidation changes from one

    nanocrystal to another. Such change within a given synthesis mode could explain the change in

    morphology obtained after oxidation and explain the formation of the two different structures, the

    yolk/shell and the hollow nanocrystals. With this in mind, the different structures obtained for Copoly-

    RM and Copoly-ORG nanocrystals respectively after oxidation at 200°C for 10 min could also be

    understood.

    From a previous study performed by the same group (7) with 7nm Copoly-RM and CoHCP-RM nanocrystals

    we known that the relative diffusion of oxygen and Co in the nanocrystals markedly changes between

    hcp single domain and polycrystalline fcc phase. Here it becomes obvious that even though the

    structure remains similar (fcc), the size of the domains significantly change the diffusion process. We

    could assume, in first approximation, that with small fcc single domain the oxygen diffuse into the Co

    lattice keeping the integrity of Co core instead of the Co diffusing outwards and creating voids within

    the nanoparticle after oxidation treatment.

    After oxidation at 260 ᵒC for 10 minutes (Figure 11g and 11h) the majority of nanoparticles show a

    hollow structure, with some elements of yolk/shell particle with a small core. The assembly on the

    TEM grid was uneven and no conclusions about self-assembly could be made. The average

    nanocrystal (11.4 ± 0.92 nm with a distribution of 8.1%) increases compare that what observed under

  • 16

    oxidation at 200°C to the size after oxidation at 260°C, as expected. The void has a slight square shape

    compared to the circular nanoparticles before oxidation and this behavior can also be seen when

    viewing the particles after oxidation at 200°C. The corresponding ED pattern (Figure 11i) shows the

    presence of Co3O4 which means the elevated temperature causes the structure to transform to a spinel

    structure instead of the rock salt structure of CoO formed after oxidation at 200°C.

    4.1.4. Single domain ε phase Co nanocrystals produced from organometallic synthesis (Coε-ORG)

    Figure 12a and 12b show single domain ε phase Co nanocrystals (Coε-ORG) produced from

    organometallic synthesis deposited on a TEM grid with an average size of 8.1 ± 0.58 nm and with a

    size distribution of 7.1% (inset Figure 12a). The large and lighter monolayers are seen with some

    small areas being double or triple layers. From the HRTEM image in Figure 13a, it is clear that the

    nanoparticle consists of one single crystalline domain displaying the ε-Co (221) lattice plane. The

    corresponding ED pattern presents a triplet, characteristic for the ε-Co (221), (310) and (311) lattice

    planes.

    Figure 12. Single domain ε phase Co nanocrystals (Coε-ORG) produced from organometallic synthesis, 8 nm. (a, b)

    TEM images before oxidation, (d, e) after oxidation at 200˚C for 10 min, (g, h) after oxidation at 260˚C for 10 min. (c,

    f, i) Corresponding ED pattern before and after the two oxidation temperatures. T marked on the ED pattern in (c)

    refers to the triplet rings [(221), (310), and (311)] of the ε lattice. The inset in (a) is a histograms displaying the particle

    diameter and the histogram insets in (d, g) displays the particle diameter and the diameter of the void.

    Like the others Co nanocrystals, the sample was submitted to an exposure of oxygen in the same

    experimental conditions as described above (200˚C; 10 min). The TEM images obtained after the

    oxidation process shows formation of hollow nanocrystals (Figure 12d and 12e). The mean diameter

    of the ε-Co nanoparticles increases compared to that obtained before oxidation and is found equal to

    9.3 ± 0.75 nm with an 8.0% distribution. The increase in size is probably due to the outwards

    migration of Co-particles during oxidation which is what also creates the central void, with an average

    diameter of 3.1 ± 0.62 nm and a distribution of 20.2%. The two measurements reveal that the shell

    thickness is 3.1 nm. The hollow CoO exist in a large quantity and only a few single core/shell

  • 17

    structures are obtained within the nanocrystal monolayer and. Some areas show particles with larger

    voids within the more ordered areas of the lattice compared to particles in the less ordered areas.

    Because of the lower order, the sample is in contact with a higher concentration of oxygen which

    makes the inward diffusion of oxygen faster and therefor the internal void, caused by outward

    diffusion of Co atoms, smaller. The HRTEM image (Figure 13b) shows that the shape of the Coε-ORG

    nanoparticles has become more irregular after oxidation at 200°C for 10 min compared to before

    oxidation (Figure 13a). The oxide shell shows regular lattice planes with a distance of 2.46 Å that

    correspond to the CoO (111) planes with two different directions. This means that the CoO shell is

    polycrystalline with two different crystalline domains. The corresponding image from the ED pattern

    (Figure 12f) shows a clear pattern from CoO, which proves the results visualized on the TEM and

    HRTEM images that a hollow CoO particle is created after oxidation at 200°C for 10 minutes.

    Figure 13. HRTEM images from single domain ε phase Co nanocrystals (Coε-ORG) produced from organometallic

    synthesis 8nm. (a) Before oxidation, single crystalline nanocrystal displaying the lattice plane from ε-Co. (b) After

    oxidation at 200˚C for 10 min. Hollow particle showing the lattice planes for the CoO shell.

    On increasing the temperature to 260°C (with 10 min exposure), hollow particles are obtained in the

    TEM grid, like after the oxidation at 200°C and this can be seen in Figure 12g and 12h. The size of the

    nanoparticle is 9.3 ± 0.69 nm, with a 7.4% distribution (inset Figure 12g), while the void has a

    diameter of 2.5 ± 0.66 nm, with a distribution of 26.7%.This information was used to calculate the

    shell thickness to 3.4 nm. This is slightly larger than the oxide thickness after oxidation at the lower

    temperature (3.1 nm). The corresponding ED pattern (Figure 12i) reveals a change in the structure

    after the change in oxidation temperature as the oxide shell is now built up by the spinal structured

    Co3O4 instead of the cubic CoO. The distance between the lattice planes is larger for the spinal

    structures and this explains the thicker shell after oxidation at the higher temperature.

    The ε phase of Co is not as stable as the two other structures. It is more easily oxidized by the electron

    beam in the TEM and it is likely that is also oxidizes easier during the oxidation treatment. The

    formation of the hollow structure leaves no trace of a core as the 8 nm samples for both CoHCP-RM and

    Copoly-ORG.

  • 18

    4.2. Size dependency

    The results from the three sizes 4, 6 and 8 nm are presented for polycrystalline nanocrystals produced

    from reverse micelle approach (Copoly-RM) and single domain hcp nanocrystals produced by annealing

    Co nanoparticles from reverse micelle approach (CoHCP-RM). Both samples, in the three sizes, are

    presented after the two oxidation treatments 200°C for 10 minutes and 260°C for 10 minutes.

    4.2.1. Polycrystalline nanocrystals produced from reverse micelles (Copoly-RM)

    A comparison between the three different sizes, 4, 6 and 8 nm, of Copoly-RM nanocrystals after

    oxidation at 200°C is presented in Figure 14. The 4 nm particles have become fully oxidized CoO

    (Figure 14a), the 6 nm particles have formed both fully oxidized CoO particles and core/shell Copoly-

    RM/CoO with a small core (Figure 14b), while the 8 nm particles almost only display a core/shell

    Copoly-RM/CoO structure (Figure 14c). The larger the particle is before oxidation, the larger the

    resulting core becomes after oxidation. A more highly ordered area promotes the formation of a core

    shell particle for a larger particle size, but as the particle size decreases all nanocrystals show complete

    oxidation independently of the degree of order of the assembly. The crystalline structure of the CoO,

    in any nanocrystal size, is similar to that of the original fcc nanoparticles with a rather low

    crystallinity. The HRTEM images from the 4 and 8 nm sample are presented in the same figure

    (Figure 14d and 14e). From the images, the smaller nanoparticles can be detected as fully oxidized

    CoO with a very low crystallinity and the 8 nm nanoparticles as core/shell nanoparticles with a

    polycrystalline CoO shell and a core with very low crystallinity.

    Figure 14. Polycrystalline nanocrystals produced from reverse micelles (Copoly-RM) after oxidation at 200°C for 10 min.

    TEM from three different sizes and HRTEM from the sizes 4 and 8 nm, (a) TEM 4 nm, (b) TEM 6 nm, (c) TEM 8 nm,

    (d) HRTEM 4 nm, (e) HRTEM 8 nm.

    The same comparison as described directly above is performed on the three sizes of Copoly-RM

    nanocrystals after the other investigated oxidation treatment at 260°C for 10 minutes and is presented

    in Figure 15. Both the TEM images and the corresponding ED pattern for the three samples are

    showed. The 4 nm nanoparticles have coalesced and also became fully oxidized which can be seen in

  • 19

    Figure 15a. The corresponding ED pattern confirms a CoO structure, but as the native Copoly-RM had

    very low crystallinity, the formed oxide has the same feature which can be seen by the unclear ED

    pattern. The 6 nm nanocrystals have also formed fully oxidized structures (Figure 15c) but the

    corresponding ED pattern (Figure 15d) reveals an uncertain oxidation state and it cannot be concluded

    if the oxide is the cubic CoO or the spinal Co3O4. One reason for this behavior can be that the 6 nm

    nanoparticle is in its transition state between the two oxidation states, but for a more elaborated

    explanation an investigation of HRTEM images would have to be made. The 8 nm nanoparticles show

    a mixture between fully oxidized and core/shell Co/Co3O4 nanoparticles (Figure 15e and 15f), the

    nanoparticles are slightly coalesced but it is less severe than for the smaller sizes.

    Figure 15. Polycrystalline nanocrystals produced from reverse micelles (Copoly-RM) after oxidation at 260°C for 10

    min. TEM images and ED patterns from three different sizes, (a) TEM 4 nm, (b) ED 4 nm, (c) TEM 6 nm, (d) ED 6

    nm, (e) TEM 8 nm, (f) ED 8 nm.

    The oxidation behavior for Copoly-RM has a clear size dependency. After both oxidation temperatures

    (200°C and 260°C) the 8 nm particles generate core/shell particles, however the oxide structure

    differs. When the particle size is decreased to 6 nm or below, the oxidation results in a fully oxidized

    structure, independently of the oxidation temperature. An interesting size dependency occurs after the

    oxidation at 260°C where the size of the nanoparticles affects the obtained oxide structure. The larger

    8 nm nanocrystals results in the spinal Co3O4 and the smaller 4 nm nanocrystal in the cubic CoO

    structure. The 6 nm nanoparticle obtains a structure which cannot be decided to be one of the

    described oxides after measuring the ED pattern, and a possible solution for this could be that the 6 nm

    size is in its transition state in between the two possible oxide structures.

  • 20

    4.2.2. Single domain hcp nanocrystals produced by annealing Co nanoparticles from reverse micelle approach (CoHCP-RM)

    A comparison between CoHCP-RM nanocrystals differing by their average size before oxidation (4nm,

    6nm and 8 nm) and oxidized in similar experimental conditions as above (200°C; 10mn) was

    performed. The resulting TEM images displays that the 4 nm CoHCP-RM nanocrystals (Figure 16a)

    obtains a hollow single crystalline CoO structure with the HRTEM image (Figure 16d) showing the

    CoO (200) lattice plane.

    On increasing the CoHCP-RM nanocrystal size to 6 nm, hollow nanocrystals with both CoO single

    domain and poly crystalline phases are observed (Figure 16b, 16e and 16f). Figure 16e shows the

    formation of single domain hollow CoO nanocrystals with presence of the (111) planes, whereas

    Figure 16f shows a hollow structure with a poly crystalline CoO lattice where the CoO (200) lattice

    planes are oriented in different directions. The hollow void for both 4 and 6 nm is smaller than the

    original Co nanoparticle indicating that the diffusion mechanism is not only outward diffusion of

    cobalt, as this would have generated a void with equal size of the original particle. The smaller central

    void can be due to some inward transportation of oxygen or due to inward relaxation of the void

    through collection of the vacancies at the interface of cobalt and cobalt oxide. (8)

    As already mentioned in chapter 4.1.2, 8 nm CoHCP-RM nanocrystals are characterized by core/shell

    Co/CoO structure with a large core, indicating that oxidation at 200°C for 10 min is not enough to

    make a nanoparticle of that size into a hollow structure.

    Figure 16. TEM and HRTEM images from the three different sizes of single domain hcp nanocrystals produced by

    annealing Co nanoparticles from reverse micelle approach (CoHCP-RM) after oxidation at 200°C for 10 min. (a) TEM 4

    nm nanocrystal, (b) TEM 6 nm nanocrystal, (c) TEM 8 nm nanocrystal, (d) HRTEM 4 nm, single crystalline hollow

    CoO particle, (e) HRTEM 6 nm, Single crystalline hollow CoO particle , (f) HRTEM 6 nm, polycrystalline hollow

    CoO particle, (g) HRTEM 8 nm, core/shell Co/CoO with large single crystalline core.

    After increasing the oxidation temperature to 260°C for 10 minutes the particles of all the sizes have

    started to coalescence. The TEM image from the 4 nm nanoparticle (Figure 17a) shows fully oxidized

    nanocrystals. The corresponding ED pattern (Figure 17b) reveals that the oxide has the cubic CoO

    structure. The 6 nm nanoparticles (Figure 17c and 17d) looks fully oxidized at the first look, but as the

    TEM image is more carefully studies some small hollow structures can be found. According to the ED

    pattern the oxide structure is in this case also the cubic CoO. The TEM image and ED pattern for the

    8 nm nanoparticles (Figure 17e and 17f) presents a hollow Co3O4 with a small void. Smaller particles

    have a higher surface to volume ratio and thereby a higher surface energy, this mean that a hollow

  • 21

    particle is less stable the smaller the particle is and could explain why the 8 nm nanoparticles result in

    a hollow particle while the 6 nm nanoparticles rarely presents a void and the 4 nm nanoparticles are all

    fully oxidized.

    Figure 17. TEM images from three different sizes of single domain hcp nanocrystals produced by annealing Co

    nanoparticles from reverse micelle approach (CoHCP-RM) after oxidation at 260°C for 10 min. TEM images and ED

    patterns from the three different sizes. (a) TEM 4 nm, (b) ED 4 nm (c) TEM 6 nm, (d) ED 6 nm, (e) TEM 8 nm, (f) ED

    8 nm.

    A clear size dependency can be seen among the CoHCP-RM nanoparticles after both oxidation

    temperatures. The lower temperature (200°C) results in the same oxide structure (CoO) as the

    oxidation for all samples at 200°C for 10 min for all sizes however the structure of the resulting

    nanoparticle differs. The 4 nm nanocrystal obtains a hollow single crystalline structure while the

    polycrystalline structure becomes more common for the 6 nm sample. The 8 nm nanocrystal does not

    form a hollow nanoparticle at all but a core/shell CoHCP-RM/CoO instead. The higher oxidation

    temperature (260°C) results in a different size dependent behavior. The larger the nanoparticle is the

    larger the internal central void becomes. For the 6 nm nanocrystals a void can be seen in some cases

    while the 4 nm nanocrystals presents a fully oxidized structure. Apart from this, the oxide structure

    clearly differs as well. For the two smaller sizes the oxide is built up by cubic CoO while the 8 nm

    nanocrystals results in the spinal Co3O4.

  • 22

    5. Conclusion

    The expected result of the investigation was to find a change within the oxidation of the various single

    domains, between the single- and polycrystals and between the different sizes of the Co nanocrystals.

    Poly- and single domain Co-nanoparticles have different oxidation behavior. Single crystalline Co-

    nanoparticles have a higher tendency to form hollow nanoparticles than polycrystalline Co-

    nanoparticles. This can be explained by the grain boundaries ability to annihilate the created vacancies

    without forming an internal void and thereby preventing the formation of a hollow nanoparticle in a

    polycrystalline structure. The grain boundaries also work as a fast diffusion track for oxygen and

    thereby enable the inward diffusion of oxygen to be faster than the outward diffusion of Co which

    hinders the formation of hollow nanocrystals. The concentration of grain boundaries are important for

    this behavior as the sample most near to be amorphous, polycrystalline nanocrystals produced from

    reverse micelles (Copoly-RM), never showed any sign of creating a hollow particle, but the other

    polycrystalline sample, Polycrystalline fcc nanocrystals produced from organometallic synthesis

    (Copoly-ORG), formed a yolk shell structure.

    The difference in oxidation behavior is also clear between the two single crystalline and between the

    two poly crystalline structures. As mentioned above the concentration of grain boundaries and thereby

    the size of the fcc domains makes up a clear different between the two polycrystalline structures. The

    difference within the Copoly-ORG sample and the different composition if the fcc domains can be the

    reason for the two different results after the oxidation treatment at 200°C, the hollow CoO and the

    yolk/shell Copoly-ORG/CoO. Regarding the single crystalline samples the hcp structure (CoHCP-RM) is

    more stable than the ε structure (Coε-ORG). This means that the cobalt atoms in the ε-lattice has a higher

    mobility when heated and therefor an easier way for diffusion. This enables the structure to be more

    readily transformed into a hollow structure since the formation of a hollow structure requires the

    outward diffusion of Co to be faster than the inward diffusion of oxygen.

    Isolated particles have in general a fast inward diffusion of oxygen as there is no protection given by

    surrounding particles. When particles are ordered in a 2D hexagonal network the neighboring particles

    slows down the rate of the inward oxygen diffusion and thereby promoted the outwards diffusion of

    Co. Isolated nanocrystals therefor have a higher probability of being fully oxidized than ordered

    nanocrystals. The oxidation behavior for the different crystal structures differ, but for all samples

    where isolated nanoparticles were investigated, the isolated particle was closer to being fully oxidized

    than the ordered nanoparticles in the same sample.

    The size of the nanocrystal as well as the oxidation temperature (with a fixed oxidation time of 10

    min) has an effect on which oxide (cubic CoO or spinal Co3O4) is obtained after oxidation. Oxidation

    at 200°C always generates the cubic oxide for the examined structures while the structure generated

    after oxidation at 260°C depends on the size and nanocrystallinity of the nanoparticle. For the single

    crystalline CoHCP-RM the spinal structure is obtained after oxidation of the 8 nm sample at the higher

    temperature while the cubic oxide is obtained for the 4 and 6 nm samples after the same treatment.

    The higher oxidation temperature (260°C) affects the behavior differently than the lower oxidation

    temperature (200°C) for the smallest single crystalline CoHCP-RM nanoparticles. The inward relaxation

    seems to be made easier by the higher temperature, as the 4 nm becomes a hollow CoO for oxidation

    at 200°C while it becomes fully oxidized CoO after oxidation at 260°C.

  • 23

    5.1. Future work

    Concluding the thesis will be a few suggestions for future work within the area. A more complete

    HRTEM study would be of interest, but as each sample takes a long time, the HRTEM images for the

    samples oxidized at 200°C for 10 minutes were prioritized during the project. However it would give a

    more detailed comparison between the two oxidation temperatures if the results were obtained for the

    samples oxidized at 260°C for 10 minutes as well.

  • 24

    References

    1. J.R. Davis. ASM Speciality Handbook: Nickel, Cobalt, and Their Alloys. pp.349-406. ASM

    international, (2000). ISBN 0-87170-685-7.

    2. J.D. Donaldson, D. Beyersmann. Cobalt and Cobalt Compounds. pp. 429-464. Ullmann's

    Encyclopedia of Industrial Chemistry, (2005). Vol. 9. DOI: 10.1002 14356007.

    3. H.C.Zeng. Synthetic architechture of interior space for inorganic nanostructures. Journal of

    Materials Chemistry. 16, (2006), 649-662.

    4. Q.Zhang, J.Xie, Y.Yu, J.Yang, J.Y.Lee. Tuningt the Crystallinity of Au Nanoparticles. Small. 6,

    (2010), 523-527. DOI: 10.1002/smll.200902033.

    5. Y.Tang, M.Ouyang. Tailoring properties and functionalities of metal nanoparticles through

    crystallinity engineering. Nature Materials. 6, (2007), 754-759. DOI: 10.1038/nmat1982.

    6. S. Sun, C.B. Murray. Synthesis of monodisperse cobalt nanocrystals and their assembly into

    magnetic superlattices (invited). Journal of Applied physics. 85, (1999), 4325 - 4330.

    7. Z.Yang, I.Lisiecki, M.Walls, M.P.Pileni. Nanocrystallinity and the Ordering of Nanoparticles in

    Two-Dimensional Superlattices: Controlled Formation of Either Core/Shell (Co/CoO) or Hollow CoO

    Nanocrystals. ACS NANO. 7, (2013), 1342–1350. DOI: 10.1021/nn304922s.

    8. Y.D.Yin, R.M.Rioux, C.K. Erdonmez, S.Hughes, G.A.Somorjai, A.P.Alivisatos. Formation of

    hollow nanocrystals through the nanoscale kirkendall effect. Science. 304, (2004), 711-714.

    9. F. Aldinger. Controlled porosity by an extreme Kirkendall effect. Acta Metallurgica. 22, (1947),

    923-928.

    10. L. Cademartiri, G.A. Ozin. Concepts of Nanochemistry. Weinheim : WILEY-VCH Verlag

    GmbH & Co. KGaA, (2009). ISBN: 978 3 527 32597 9.

    11. H.J. Fan, U. Gösele, M. Zacharias. Formation of Nanotubes and Hollow Nanoparticles Based on

    Kirkendall and Diffusion Processes: A Review. Small. 3, (2007), 1660-1671.

    12. V.A. de la Peña O’Shea et al. Electronic and magnetic structure of bulk cobalt: The α, β, and ε-

    phases from density functional theory calculations. The Journal of Chemical Physics. 133, (2010),

    024701(1)-(8).

    13. M. Cavalier, M. Walls, I. Lisiecki, M. P. Pileni. How Can the Nanocrystallinity of 7 nm

    Spherical Co Nanoparticles Dispersed in Solution Be Improved? Langmuir. 27, (2011), 5014–5020.

    Avaliable from: dx.doi.org/10.1021/la1049887 .

    14. I. Lisiecki, M. Walls, D. Parker, M.P. Pileni. 2D Self-Organisation of Core/Shell Cohcp/Co

    Nanocrystals. Langmuir. 24, (2008), 4295 - 4299.

    15. P.Patnaik. Handbook of Inorganic Chemical Compounds. s.l. : mcgraw-Hill Professional, (2002).

    ISBN: 0070494398.

    16. S.C.Petitto, E.M.Marsh,G.A.Carson, M.A.Langell. Cobalt oxide surface chemistry: The

    interaction of CoO(1 0 0), Co3O4 (1 1 0), and Co3O4 (1 1 1) with oxygen and water. Journal of

    Molecular Catalysis A: Chemical. 128, (2008), 49-58.

    17. J. Shi. Steric Stabilization. The Ohio State University. [Online] 08 29, 2002. [Cited: 02 16, 2013.]

    http://www.matsceng.ohio-state.edu/ims/LR_Stericstablization.pdf.

    18. Sigma-Aldrich. Sigma-Aldrich: Oleic Acid. [Online] 2013. [Cited: 04 25, 2013.]

    http://www.sigmaaldrich.com.

    19. L.E. Smart, E.A. Moore. SOLID STATE CHEMISTRY An Introduction. Boca Raton : Taylor &

    Francis group, LLC, (2005). ISBN: 978 0 7487 7516 3.

    20. F. Krumeich. Electron Diffraction (ED). Electron Microscopy ETH Zürich. [Online] 05 14, 2012.

    [Cited: 01 25, 2013.] http://www.microscopy.ethz.ch/ED-1.htm.